You are on page 1of 18

International Journal of Numerical Methods for Heat & Fluid Flow

A numerical model for the thermocapillary flow and heat transfer in a thin liquid film
on a microstructured wall
A. Alexeev T. Gambaryan-Roisman P. Stephan

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

Article information:
To cite this document:
A. Alexeev T. Gambaryan-Roisman P. Stephan, (2007),"A numerical model for the thermocapillary flow and
heat transfer in a thin liquid film on a microstructured wall", International Journal of Numerical Methods for
Heat & Fluid Flow, Vol. 17 Iss 3 pp. 247 - 262
Permanent link to this document:
http://dx.doi.org/10.1108/09615530710730139
Downloaded on: 05 April 2016, At: 11:34 (PT)
References: this document contains references to 13 other documents.
To copy this document: permissions@emeraldinsight.com
The fulltext of this document has been downloaded 364 times since 2007*
Access to this document was granted through an Emerald subscription provided by emerald-srm:596032 []

For Authors
If you would like to write for this, or any other Emerald publication, then please use our Emerald for
Authors service information about how to choose which publication to write for and submission guidelines
are available for all. Please visit www.emeraldinsight.com/authors for more information.

About Emerald www.emeraldinsight.com


Emerald is a global publisher linking research and practice to the benefit of society. The company
manages a portfolio of more than 290 journals and over 2,350 books and book series volumes, as well as
providing an extensive range of online products and additional customer resources and services.
Emerald is both COUNTER 4 and TRANSFER compliant. The organization is a partner of the Committee
on Publication Ethics (COPE) and also works with Portico and the LOCKSS initiative for digital archive
preservation.
*Related content and download information correct at time of download.

The current issue and full text archive of this journal is available at
www.emeraldinsight.com/0961-5539.htm

A numerical model for the


thermocapillary flow and heat
transfer in a thin liquid film on a
microstructured wall
Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

A. Alexeev
Department of Chemical Engineering, University of Pittsburgh,
Pittsburgh, Pennsylvania, USA, and

Thermocapillary
flow and heat
transfer
247
Received 1 January 2006
Accepted 9 July 2006

T. Gambaryan-Roisman and P. Stephan


Darmstadt University of Technology, Darmstadt, Germany
Abstract
Purpose This paper aims to study thermocapillarity-induced flow of thin liquid films covering
heated horizontal walls with 2D topography.
Design/methodology/approach A numerical model based on the 2D solution of heat and fluid
flow within the liquid film, the gas above the film and the structured wall is developed. The full
Navier-Stokes equations are solved and coupled with the energy equation by a finite difference
algorithm. The movable gas-liquid interface is tracked by means of the volume-of-fluid method. The
model is validated by comparison with theoretical and experimental data showing a good agreement.
Findings It is demonstrated that convective motion within a film on a structured wall exists at any
nonzero Marangoni number. The motion is caused by surface tension gradients induced by
temperature differences at the gas-liquid interface due to the spatial structure of the heated wall. These
simulations predict that the maximal flow velocity is practically independent from the film thickness,
and increases with increasing temperature difference between the wall and the surrounding gas. It is
found that an abrupt change in wall temperature causes rupture of the liquid film. The thermocapillary
convection notably enhances heat transfer in liquid films on heated structured walls.
Research limitations/implications Our solutions are restricted to the case of periodic wall
structure, and the flow is enforced to be periodic with a period equal to that of the wall.
Practical implications The reported results are useful for design of the heat transfer equipment.
Originality/value New effects in thermocapillary convection are presented and studied using a
developed numerical model.
Keywords Volume measurement, Fluid dynamics, Numerical control, Convection, Heat transfer
Paper type Research paper

Nomenclature
A
Bi
C
C~

wall structure amplitude


Biot number
color function
averaged color function

cp
d
F
Fs

specific heat capacity


wall structure period
body forces
body force due to surface tension

The authors would like to acknowledge the generous support of the German Science Foundation,
DFG, through the Emmy Noether Program. A. Alexeev is grateful to Virgil Stoica for his
assistance in the conduction of the experiments.

International Journal of Numerical


Methods for Heat & Fluid Flow
Vol. 17 No. 3, 2007
pp. 247-262
q Emerald Group Publishing Limited
0961-5539
DOI 10.1108/09615530710730139

HFF
17,3

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

248

hg
hw
Dh
*
Dh
k
M
M cr
n
Nu
p
S
t
T
DT
u
umax

gas layer thickness


wall structure height
average thickness of liquid film
minimal thickness of liquid film
interface curvature
Marangoni number
critical Marangoni number
unit normal to interface
Nusselt number
pressure
interface height function
time
temperature
temperature drop
velocity
maximal liquid velocity at interface

Greek symbols
a
thermal diffusivity

ds
l
m
r
s
sT
t
w

Dirac distribution function at


interface
thermal conductivity
dynamic viscosity
density
surface tension
temperature coefficient of surface
tension
shear stress tensor
wall structure angle

Superscripts
n
iteration level
Subscripts
0, g gas
1
liquid
i, j
computational cell index
w
wall

Introduction
Surface tension is crucial in the dynamics of thin liquid films on substrates of different
topography, which are frequently encountered in many engineering applications,
including the thermal management of electronic devices, food processing, chemical
engineering, MEMS.
Normally, the surface tension of a liquid is a decreasing function of temperature.
If the temperature varies at the gas-liquid interface, surface tension gradients cause
thermocapillary (Marangoni) flow (Colinet et al., 2001). If a thin liquid film is heated on
a planar substrate of a uniform temperature, a conducting solution exists, which
implies that the film is motionless and the free surface of the liquid is isothermal. This
solution is stable for sufficiently small temperature gradients across the liquid layer.
If the temperature gradient exceeds a critical value, the conducting solution loses
stability and convective patterns are developed (Pearson, 1958). If the substrate has a
structure on its surface, the convection prevails for any temperature difference
(Alexeev et al., 2005). It occurs due to the temperature inhomogeneity, which is imposed
on the interface by the spatial structure of the substrate.
We develop a numerical model to describe the motion of a thin liquid film on a
heated structured wall. We deploy a finite difference algorithm to integrate the
Navier-Stokes and energy equations. To cope with the movable gas-liquid interface, we
apply the volume-of-fluid (VOF) method (Scardovelli and Zaleski, 1999). The
calculations are performed simultaneously through the whole computational domain
containing the gas and liquid regions, while the surface force at the interface is
included into the momentum balance equations via a volumetric force (Kothe and
Mjolsness, 1992). The energy equation is solved within the solid wall as well.
There are only few previous studies where the Marangoni flows in a cavity are
considered in the framework of the VOF. Sasmal and Hochstein (1994) calculated the
Marangoni convection induced by a temperature difference between the sidewalls of a
rectangular cavity. They studied heat transfer within the cavity and the effect of the
contact angle on the flow patterns. More recently, Wang (2002) applied a VOF model to

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

investigate the Marangoni convection in trapezoidal cavities. In these works, the


temperature gradient was caused by a temperature difference between the sidewalls.
Thus, the heat flux was mostly directed along the gas-liquid interface, in that way it
was justified to neglect its component normal to the interface by considering the
adiabatic condition at the free surface. In contrast, in a flow on a heated structured
wall, the heat flux is practically perpendicular to the gas-liquid interface resulting in a
strong temperature gradient in that direction. In this case, the thermocapillary force is
induced by a relatively small variation of the liquid temperature along the interface.
Hence, a very accurate calculation of the temperature is required to avoid unphysical
flow currents due to inaccuracy in the temperature gradient evaluation.
In present work, we study thermocapillary motion within a thin film of a low
volatility liquid on a heated highly thermal conductive wall with 2D microscale
topography. We consider a situation where the liquid layer covers a horizontal wall,
and its thickness is comparable with the amplitude of the wall microstructure. We
neglect the effect of gravity. Our solutions are restricted to the case of periodic wall
structure, and the flow is enforced to be periodic with a period equal to that of the wall.
Numerical model
Governing equations
The incompressible flow is governed by the continuity equation:
7 u 0;
and the Navier-Stokes equations:


u
u 7u 27p 7 t F;
r
t

where u is the velocity, r the density, p the pressure, F the body forces, and t time.
Moreover, t is the shear stress tensor given by:


m uj ui
t ij

;
3
2 xi xj
where m is the dynamic viscosity.
The equations are coupled with the energy equation given by:


T
u 7T 7 l7T;
r cp
t

where T is the temperature, cp the specific heat capacity, and l the thermal
conductivity.
To track the moving gas-liquid interface, we utilize the VOF technique (Scardovelli
and Zaleski, 1999). A color function C is introduced, which equals to 1 within the liquid
and to 0 within the gas. The color function is governed by a transport equation:

C
u 7C 0:
t

We are looking for the solutions, which are characterized by a period equal to that of
the wall structure, d (Figure 1). Thus, we impose a symmetry boundary condition at

Thermocapillary
flow and heat
transfer
249

HFF
17,3

y
hg

Gas

Liquid

250

Wall

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

Tw

hw

h*

TwT

Figure 1.
Outline of the structured
wall from the experiments

x 0 and x d=2. We also impose T T w at y 0 and T T w 2 DT at


y hw Dh * hg . Moreover, a free flow condition for the velocity u=y 0 is
applied at y hw Dh * hg .
Numerical method
The hydrodynamic equations (1)-(3) are solved with a finite difference algorithm on a
rectangular staggered grid using the projection method (Ferziger and Peric, 2002). The
projection method consists of three steps. First, the prediction velocity field u* due to
the advective and diffusive terms in equation (2) is to calculate semi-implicitly:
Du
1
1
1
u n 7Du 2 n 7 Dt 2u n 7u n n 7 t n n F n ;
Dt
r
r
r

where Du u* 2 u n and Dt t* 2 t n . Then, the Poisson equitation, which is


obtained using equation (1), is solved to calculate the pressure field:

7

1
7p n1
rn

7 u*
:
Dt

Finally, the velocity field is corrected to the time level n 1:


u n1 u*

Dt
7 u* :
rn

Using the velocity field u n1 , the color function is advected by solving:


DC
u n1 7DC 2u n1 7C n ;
Dt

where DC C n1 2 C n . The final stage of the numerical solution involves calculation


of the temperature field:

Thermocapillary
flow and heat
transfer

DT
1
u n1 7DT 2
7 l n1 7DT
Dt
cp rn1

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

2u n1 7T n

1
7 l n1 7T n :
cp rn1

10

Here, DT T n1 2 T n .
The advection terms in the r.h.s. of equations (6), (9) and (10) are solved with the
third order essentially non-oscillatory (ENO) scheme (Shu and Osher, 1988), while the
terms in the advection terms in the l.h.s. of these equations as well as the viscous and
conductivity terms are approximated with the second order finite differences. The ENO
scheme is used since it provides good tracking of the discontinuity-like interfaces.
To treat implicit parts of equations (6), (9) and (10), we utilize the approximate
factorization approach and solve the equations separately along the x- and y-directions.
The overall accuracy of our method is of the second order in space.
The multigrid technique (Wesseling, 1991) is applied to solve the Poisson equitation
for pressure (equation (7)). We use v-cycle and the number of multigrid levels K is
given by 3 2K minN x ; N y , where N x and N y are the grid size in the x- and
y-directions, respectively. The use of the multigrid technique reduces the overall
computational time by an order of magnitude as compared to the standard iterative
methods.
To impose the non-slip velocity condition at the liquid-solid boundary, we utilize the
immersed boundary approach. We set the x and y velocity components within the solid
domain at the nodes right next to the liquid-solid interface such that the linearly
interpolated velocity at the interface equals to zero. We also set at these nodes a zero
gradient for pressure, while solving equation (7).
Following the VOF approach, we calculate the values of density and viscosity used
in equations (6)-(10) as:
~ r0 C~ r1 ;
r 1 2 C

~ m0 C~ m1
m 1 2 C

11

Hereafter, the index 1 stands for the liquid, while the index 0 denotes the gas
properties. Moreover, C~ is the averaged color function (Alexeev et al., 2005).
Accurate calculation of the temperature distribution along the gas-liquid interface
is critical for a correct modeling of thermocapillary driven flows. When the heat flux is
directed across the interface, the difficulty arises due to the discontinuity of properties
of the fluids across the interface. In this case, the cell average values cannot provide a
satisfactory description for the fluid properties at the interface. In particular, our
simulations show that the use of an averaging, either algebraic or geometric, for the
calculation of l causes spurious currents in the fluids.
Mehdi-Nejad et al. (2005) have recently developed an approach for a more accurate
calculation of the convection terms in the energy equation. They successfully applied
this approach to study heat transfer in molten tin drops during their fall. In the case of
thermocapillary driven flows on heated walls, however, the heat flux across the
interface is typically dominated by the diffusive rather than convective terms.
We, therefore, propose a simple approach to calculate the temperature flux across the
interface. Consider an interface that is located at the cell i; j such that the upper and
bottom boundaries of the cell are along the interface (Figure 2). To resolve the diffusive

251

HFF
17,3

y
Ci,j+1=0
Ti,j+1

Dy

Ti,j+1
i,j+1

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

252

T 0i,j

Dy0

T ii,j
Ti,j

Dy1

Figure 2.
Schematic diagram of
computational cells near
the interface and an
approximation of the
temperature distribution
across the interface

T 1i,j

i,j
Ci,j-1=1
Dy

interface
Ti,j-1

Ti,j-1

i,j1

terms in equation (10), we approximate the heat fluxes across the bottom and upper
boundaries of i; j as:
f2
y 2l1

T 1i;j 2 T i;j21
T i;j1 2 T 0i;j

and
f

2
l
;
0
y
Dy Dy 1
Dy Dy 0

12

respectively. Here, Dy 1 C i; j Dy, Dy 0 1 2 C i; j Dy and Dy is the computational grid


step. The temperatures T 0i; j and T 1i; j are calculated using two conditions: continuity of
the temperature at the interface:

 Dy 2Dy 0
T ii;j T i;j1 T 0i;j 2 T i;j1
;
Dy Dy 0

13a


 Dy 2Dy 1
:
T ii;j T i;j21 T 1i;j 2 T i;j21
Dy Dy 1

13b

and energy conservation within the cell i; j:

rcp T i;j rcp 0 T 0i;j 1 2 C i;j rcp 1 T 1i;j C i;j ;

14

rcp 1 2 C i;j rcp 0 C i;j rcp 1 :

15

where:

T 0i;j

T 1i;j

and
can be readily calculated. These
Combining equations (13)-(15),
temperatures are also used to evaluate the temperature gradients within the liquid near
the gas-liquid interface needed to calculate the thermocapillary force acting at the
interface.
To estimate the heat flux component, which is directed along the interface at the cell
i; j, an algebraic averaging for l is used (equation (10)). At the liquid-solid interface,

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

where we do not need an accurate value of the temperature gradients, the geometric
averaging is used to assess the vertical heat flux, while the algebraic averaging is
applied in the horizontal direction.
To illustrate the method for the temperature calculation, we solve equation (10) for a
test problem in which the interface between two motionless fluids u 0 is slightly
inclined and the heat flux is due to a temperature difference between the upper and
lower walls (Figure 3(a)). In Figure 3(b) and (c), we present the x and y components
of the temperature gradient along the interface, respectively. We are interested in the
temperature gradient since the thermocapillary force is directly proportional to its
magnitude. To compare with our approach, we also include the results for the
temperature gradients calculated with the algebraic and geometric averaging of the
thermal conductivity within the numerical cells at the interface.
One can expect for the considered problem (Figure 3(a)) that the interface
temperature changes monotonically, meaning that the magnitude of the temperature
gradient may not oscillate along the interface. Nevertheless, both the algebraic and
geometric averaging results in strong oscillations of the temperature gradient with a
period which is correlated with the numerical grid spacing. These oscillations
eventually cause spurious currents along the interface induced by the unphysical
variations in the thermocapillary force. In contrast, our approach gives a much better
approximation of the gradients along the interface. Although there is still some noise in
T=x due to the discretization, it can be reduced by applying an appropriate
smoothing. Our simulations, however, show that this noise practically does not affect
the results.
To include the effect of surface tension into the momentum equations, we adopt the
continuum surface force approach (Kothe and Mjolsness, 1992). The body force due to
surface tension is given by:

Fs

2rds
skn 1 2 n^n7s;
r0 r1

16

~ is the Dirac distribution function at the


where s is the surface tension, ds j7Cj
interface, k is the curvature of the interface, and n is the unit normal to the interface.
Moreover, 7s sT 7T, where sT is the temperature coefficient of surface tension.
To obtain 7T at the gas-liquid interface, we calculate the temperature gradients at
the cells near the interface, which are filled with the liquid, and then extrapolate
the gradients to the interface. To assess the unit normal n and the curvature of the
interface k, we utilize a reconstruction algorithm (Sussman, 2003), which is based on
reconstructing the height function S directly from the color function C.
Although the ENO scheme, which is used to calculate C provides good tracking of
the interface, it causes some numerical foam around the interface, which can
be accumulated during long time calculations. We, therefore, restore C near the
interface in such a way that the height functions S and the normal n remain
unchanged, while the foam is eliminated. In fact, this procedure breaks the global
mass conservation. Our simulations show, however, that the change of the mass is
rather small and usually does not exceed 102 3 of its initial value even for long
calculations.

Thermocapillary
flow and heat
transfer
253

dT/dx=0

Gas

254

TwT
hmax

Interface
hmin

dT/dx=0

HFF
17,3

Liquid
d

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

Tw
(a)
0.001

dT/dx d/T

0.0005
0
0.0005
0.001

0.2

0.4

0.6

0.8

x/d
(b)
0.06

dT/dy d/T

0.05
0.04
0.03
0.02
0.01
0

0.2

0.4

0.6

0.8

x/d
(c)

Figure 3.
Test problem for the
solution of equation (4) for
an inclined interface
between two domains of
an equal average
thickness

Notes: The domains have thermal properties of a gas (air) and a liquid (water).
Grid size is 96 96; domain size is h = 0.5 (hmin + hmax) =1 mm,
h=hmax hmin = 0.1 mm (~ 5y), d = 5mm; T = 10K: (a) schematic of
the test problem; (b) and (c), respectively, represent the horizontal and vertical
components of the temperature gradient along the interface calculated for
different approximations for l . The solid lines show the approximation of
equations (12)-(15), the dotted and dashed lines are for the algebraic and
geometric averaging, respectively

Computational parameters
We carry out the simulations for two liquids, which are water and silicon oil, while the
gas is air. Their properties are chosen at Tw 238C. The calculations are performed
for two types of wall structures. The first wall (Figure 1) corresponds to the
experiments reported in Alexeev et al. (2005) with d 1 mm, hw 0.5 mm and
f 308. The second wall is given by:

255

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)



2px
yw A 1 2 cos
;
d

17

where A hw =2 is the wall structure amplitude. Properties of the wall material are
those of copper. We also set in all our calculations hg 0.4 mm.
We perform the calculations for a half of the groove. The computational domain is
0 # x # d=2 and 0 # y # hw Dh * hg (Figure 1). Our rectangular computational
grid usually consists of 96 96 cells. To test the grid quality its density was
increased, indicating that an increase in grid density practically does not affect the
solution.
The calculations are started with zero initial velocities and continued up to the
moment when a steady state solution is obtained.
Results and discussion
Model validation
We first consider thermocapillary convection in a rectangular cavity due to a
temperature difference between the sidewalls. In the limit of thin film within a wide
cavity, this problem can be solved analytically (Levich, 1962) and, therefore, can serve
as a test case for our numerical model. Figure 4(a) and (b) shows the velocity and
temperature distributions within the cavity for silicon oil and water, respectively.
As expected, the thermocapillary force induces vortexes within the fluids in which the
flow near the interface is directed toward the wall having lower temperature. Note that
the isotherms within the silicon oil (Figure 4(a)) are practically vertical that
corresponds to the conducting solution, while in Figure 4(b) for water, they are
considerably distorted by the flow. This difference arises because water has a lower
Prandtl number as compared to the silicon oil.
Figure 5 shows the pressure distribution along the gas-liquid interface the solutions
shown in Figure 4 as well as the theoretical prediction (Levich, 1962). As seen, there is
good agreement between the numerical solutions and the theory. Some discrepancy
near the sidewalls can be attributed to the fact that the free surfaces are deformed by
the flow, while the theory neglects this effect.
To verify our numerical model for the case when the thermocapillary
convection is driven by a vertical temperature gradient, we first consider a
rectangular cavity and estimate the critical Marangoni number M cr corresponding
to the onset of thermocapillary convection (Colinet et al., 2001). The Marangoni
number is given by:
M

sT DT 1 Dh
;
r1 n1 a 1

Thermocapillary
flow and heat
transfer

18

0.8

0.2

0.4

0.6

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

256

where DT 1 DTBi1 Bi 21 is the temperature drop over the liquid, a


the thermal diffusivity, Dh the average thickness of the liquid film and
Bi l0 Dh=l1 hg the Biot number. In our simulations, we found good agreement
with the linear theory (Colinet et al., 2001). Namely, for M less than the theoretical
value of M cr (Colinet et al., 2001), an initial disturbance introduced into the
conducting temperature distribution decays and the fluid flow stops after a
transient, while for M . M cr , steady vortexes develop.
For more thorough model validation in the case of heated grooved walls, we
conducted experiments with a thin film of silicon oil (5cSt) on a wall with a structure
shown in Figure 1. In the experiments, we measured temperature of the film as well as

0.4

HFF
17,3

Gas

0.8

0.6

0.4

0.2

0.2

y/d

0.3

Liquid
0.1
0.8

0.6

0.8

0.8

0.6

0.8

0.6

0.8

0.4

0.6

0.2

0.6

0.4

0.2

0.2

0.4

0.4

x/d
(a)

Gas

0.2

0.4

0.2

y/d

0.3

Liquid

0.2

0.4

0.2

Figure 4.
Velocity and temperature
distributions within a
liquid film in a rectangular
cavity due to a
thermocapillary flow
induced by a temperature
difference between the
sidewalls

0.1

0.4

0.6
0.8
1
x/d
(b)
Notes: T = 1K, h = 0.1 mm, d = 0.5mm. The liquids are: (a) silicon oil; and
(b) water. The arrows show the velocity field. The thin lines represent isotherms.
The temperature is normalized as (T Tw) /T

Thermocapillary
flow and heat
transfer

5
CFD Silicon oil
pd/(T)

2.5

CFD Water
Theory

257

0.2

0.4

0.6

0.8

x/d
Notes: T = 1K, h = 0.1 mm, d = 0.5 mm

Figure 5.
Pressure distribution
along the gas-liquid
interface in a rectangular
cavity due to a
temperature difference
between the sidewalls

its maximal velocity. A detailed description of the experimental setup and the
procedure can be found elsewhere (Alexeev et al., 2005).
Figure 6 shows a numerically calculated flow pattern for M < 1 ! M cr .
The simulation predicts the formation of a vortex, in which the liquid near the
free surface moves toward the groove trough at x 0. This convection is induced by
the thermocapillary force due to a temperature gradient along the gas-liquid interface,
which is originated from the topography of the heated wall. Moreover, in agreement
with the experiments (Alexeev et al., 2005), our simulations predict that the convection
in films on structured walls arises for any temperature difference across the film.
In Figure 7(a), we compare the experimentally measured maximal values of the liquid
velocity at the gas-liquid interface, umax , with the predictions of the numerical model.
1
0.2

0.2

Gas

0.8

0.4
0.5
0.6

0.6

0.4
0.5

0.6

0.7
0.8

0.7
0.8 .9
0

y/d

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

2.5

0.4
0.9

Liquid

0.2

Wall
0
0

0.2

0.4
x/d

Notes: h* = 0.1mm, T = 1.5 K, M 1. The arrows show the velocity field.


The thin lines represent isotherms. The temperature is normalized as
(T Tw + T) /T

Figure 6.
Velocity and temperature
distributions in a silicon
oil film on a structured
wall (Figure 1)

HFF
17,3

umax, mm/s

258

M=Mcr

Experiment
Numerical results

1.5
1

0
0

10

20
TwT, K
(a)

30

40

2
Experiment
Numerical results:
with convection
conduction only

1.5
TwTint, K

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

0.5

Figure 7.
(a) Maximal velocity of the
liquid as a function of
wall temperature;
(b) temperature drop
over the liquid layer as
a function of wall
temperature

M=Mcr

1
0.5
0
0

10

20

30
TwT, K

40

50

60

(b)
Notes: Silicon oil, h* = 0.5 mm

One can see that there is reasonable agreement between the model and the experiments,
although almost everywhere the model prediction is slightly below the experimental
data. When the temperature exceeds the value corresponding to M cr , the velocity
increases much faster than that for the lower temperatures. It is an effect of the
convective motion within the liquid layer on heat transfer. Indeed, the circulations in
the layer bring hotter liquid from the hot wall to the interface, thus increasing the
temperature gradient and the thermocapillary force that in turn increases the velocity.
Figure 7(b) shows the measured and calculated temperature drop across the liquid
film as a function of the wall temperature. To demonstrate the effect of the convection on
heat transfer, the temperature drop is also calculated when the convective terms are
omitted from equation (4). In this case, the temperature drop increases linearly with T w ,
while in the calculations with the convection terms and in the experiments, the
temperature drop declines from the straight line for M . M cr . This result suggests that
there is an increase in heat transfer due to the Marangoni convection within the liquid.
We conclude that the overall agreement between our calculations and the theory
and experiments is rather convincing, and our numerical model may be applied to
study thermocapillary flows within thin films on structured walls. In what follows, we
present some numerical solutions, which are characteristic of such flows.

Thermocapillary
flow and heat
transfer
259

umaxh / 1

100

10

A=0.25mm
A=0.125mm
A=0.25mm
hw=0.5mm

1
10

100

1000

d=2mm
d=2mm
d=1mm
d=1mm

10000

M
(a)
1000

umaxh / 1

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

Flow velocities
Figure 8 shows the dependence of umax on the Marangoni number. In Figure 8(a), the
temperature drop is fixed at DT 10 K and M changes due to Dh * . In Figure 8(b), M
increases due to the increasing temperature difference, while the other parameters are
constant.
As seen in Figure 8(a), all the data collapse into a single curve. For M . 100,
umax , a1 Dh 21 M 1=2 . Taking into account that DT const, we obtain that M
changes as Dh 2 , and, therefore, umax , const. It means that for larger M , umax
practically does not depend on the thickness of the liquid layer.
The temperature difference DT, however, does modify the velocity as shown in
Figure 8(b). Note that umax grows exponentially with increasing M . It is interesting
that for different Dh * , umax follows a common curve. Our calculations show that for
Dh * const, u , DT 2=3 .

100

h*/d=0.025

h*/d=0.1

h*/d=0.05
h*/d=0.2
h*/d=0.5

h*/d=0.2
h*/d=0.4
h*/d=0.6

10

1
10

100

1000
M
(b)

10000

100000

Notes: The liquid is water. The empty markers denote calculations for a sinusoidal
wall profile, while the full markers stand for calculations with a wall shown
in Figure 1: (a) T = 10K; (b) A = 0.25 mm, d = 2 mm

Figure 8.
Maximal velocity of liquid
vs the Marangoni number

HFF
17,3

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

260

Temperature transient
The simulations show that when the wall temperature increases suddenly, the liquid
film may be ruptured as shown in Figure 9. In this simulation, uniform initial
temperature equal to T w 2 DT is imposed. At t 0, T w is applied at y 0. The
rupture is caused by a large temperature gradient along the free surface, which is
induced when the thermal boundary layer initially formed along the liquid-wall
interface reaches the gas-liquid interface near the groove crest.
We note that for the parameters in Figure 9, a steady state solution may be obtained
either by a steady increase in the wall temperature or when the temperature change
takes place at gas above the film. In the latter case, a thermal boundary layer is formed
first at the gas-liquid interface. The temperature propagates perpendicular to the free
surface, preventing the appearance of strong temperature gradients causing the
rapture.
Heat transfer
Figure 10 shows the Nusselt number as a function of Dh * for water and silicon oil.
In these calculations, we consider a wall having a structure shown in Figure 1, which is
either heated or cooled.
When DT 210 K, i.e. the wall is cooled, Nu exceeds unity by few percents for
small Dh * only, while for larger Dh * , Nu is practically equal to unity. Thus, the effect
of convection on heat transport from a cooled structured wall is relatively weak and
prevails for small Dh * only.
For a heated wall, however, our simulations predict that heat transfer can be
significantly enhanced by the Marangoni convection within the film. In this case, Nu
0.5

Gas
0.4

0.1

0.1

y/d

0.2

Liquid

3
0.40.
6
.
6
.
0 0

0.2
0.1
0
Figure 9.
Velocity and temperature
distribution in a liquid
film just before rupture
due to an abrupt increase
in wall temperature

Wall

0.1
00.02.4.3
0.6 0.8

0.1

0.4
0.6

0.1

0.3

0.2
0.3

0.2

0.3

0.4

0.5

x/d
Notes: The temperature drop T = 87K (M = 1,435) corresponds to the
minimal temperature at which a water film with h* = 0.1 mm
is ruptured on a sinusoidal wall with A = 0.25 mm, d = 2 mm.
The arrows show the velocity field.The thin lines represent
isotherms. The temperature is normalized as (T Tw + T )/T

1.7

Thermocapillary
flow and heat
transfer

T= 10K
T= 10K

1.6

Nu

1.5
1.4

261

1.3
1.2

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

1.1
1
0

0.2

0.4

0.6

0.8

1.2

Film thickness, h*/d


Notes: The liquid is water. The empty markers denote calculations for a sinusoidal
wall with A = 0.25 mm, d = 1 mm. The full markers stand for calculations with
a wall shown in Figure 1

has maximum, which occurs for about the same optimal film thickness Dh*opt for both
liquids. Thus, Dh*opt corresponds to a layer thickness, for which the effect of convection
on heat transfer is most pronounced. Our simulations show that if Dh * . Dh*opt , a
stagnant layer of liquid is formed under the vortexes attached to the gas-liquid
interface. This stagnant layer reduces the convective heat transfer from the wall,
resulting in a decrease in Nu.
We also note in Figure 10 that the sinusoidal wall provides better convective heat
transport compared to the experimental wall. Thus, an appropriate choice of the
structure can enhance heat transfer in thin liquid films on structured walls.
Summary
We study thermocapillarity-induced flow of thin liquid films covering heated
horizontal walls with 2D topography. To this end, we develop a numerical model based
on the integration of the Navier-Stokes and energy equations by a finite difference
algorithm. The mobile gas-liquid interface is tracked with the VOF technique. The
numerical model is verified by comparison with a theory and experiments showing
good agreement.
We demonstrate that convective motion within a film on a structured wall exists at
any nonzero Marangoni number. The motion is caused by surface tension gradients
induced by temperature differences at the gas-liquid interface due to the spatial
structure of the heated wall. Our simulations predict that the maximal flow velocity,
which occurs at the gas-liquid interface, is practically independent from the thickness
of the liquid layer, and increases according to a power-law with increasing DT.
It is found that an abrupt change in wall temperature causes rupture of the liquid
film near the structure crest. The rupture may occur at the same value of DT, for which
a steady state solution exists and can be obtained either by a gradual increase in wall
temperature or cooling gas above the liquid.
We show that the thermocapillary convection notably enhances heat transfer in
liquid films on heated structured walls. An optimal film thickness exists for which Nu
attains the maximal value for a specific temperature drop.

Figure 10.
Numerically calculated
Nusselt number as a
function of film thickness

HFF
17,3

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

262

References
Alexeev, A., Gambaryan-Roisman, T. and Stephan, P. (2005), Marangoni convection and heat
transfer in thin liquid films on heated walls with topography: experiments and numerical
study, Physics of Fluids, Vol. 17, pp. 062106-13.
Colinet, P., Legros, J.C. and Velarde, M.G. (2001), Nonlinear Dynamics of Surface-Tension-Driven
Instabilities, Wiley, New York, NY.
Ferziger, J.H. and Peric, M. (2002), Computational Methods for Fluid Dynamics, Springer, New
York, NY.
Kothe, D.B. and Mjolsness, R.C. (1992), RIPPLE: a new model for incompressible flows with free
surfaces, AIAA Journal, Vol. 11, pp. 2694-700.
Levich, V.G. (1962), Physicochemical Hydrodynamics, Prentice-Hall Inc., Englewood Cliffs, NJ.
Mehdi-Nejad, V., Mostaghimi, J. and Chandra, S. (2005), Modeling interfacial heat transfer from
single or multiple deforming droplets, International Journal of Computational Fluid
Dynamics, Vol. 19, pp. 105-13.
Pearson, J.R.A. (1958), On convection cells induced by surface tension, Journal of Fluid
Mechanics, Vol. 4, pp. 489-500.
Sasmal, G.P. and Hochstein, J.I. (1994), Marangoni convection with a curved and deforming free
surface in a cavity, ASME Journal of Fluids Engineering, Vol. 116, pp. 577-82.
Scardovelli, R. and Zaleski, S. (1999), Direct numerical simulation of free-surface and interfacial
flow, Annual Review of Fluid Mechanics, Vol. 31, pp. 567-603.
Shu, C.-W. and Osher, S. (1988), Efficient implementation of essentially non-oscillatory
shock-capturing schemes, Journal of Computational Physics, Vol. 77, pp. 439-71.
Sussman, M. (2003), A second order coupled level set and volume-of-fluid method for computing
growth and collapse of vapor bubbles, Journal of Computational Physics, Vol. 187,
pp. 110-36.
Wang, G. (2002), Finite element simulations of free surface flows with surface tension in
complex geometries, ASME Journal of Fluids Engineering, Vol. 124, pp. 584-94.
Wesseling, P. (1991), An Introduction to Multigrid Methods, Wiley, New York, NY.
Corresponding author
P. Stephan can be contacted at: pstephan@ttd.tu-darmstadt.de

To purchase reprints of this article please e-mail: reprints@emeraldinsight.com


Or visit our web site for further details: www.emeraldinsight.com/reprints

Downloaded by University Of Nottingham Malaysia Campus At 11:34 05 April 2016 (PT)

This article has been cited by:


1. Tatiana Gambaryan-Roisman. 2015. Modulation of Marangoni convection in liquid films. Advances in
Colloid and Interface Science 222, 319-331. [CrossRef]
2. Valeri Frumkin, Wenbin Mao, Alexander Alexeev, Alexander Oron. 2014. Creating localized-droplet train
by traveling thermal waves. Physics of Fluids 26, 082108. [CrossRef]
3. Wenbin Mao, Alexander Oron, Alexander Alexeev. 2013. Fluid transport in thin liquid films using
traveling thermal waves. Physics of Fluids 25, 072101. [CrossRef]
4. C. Ma, D. Bothe. 2013. Numerical modeling of thermocapillary two-phase flows with evaporation using
a two-scalar approach for heat transfer. Journal of Computational Physics 233, 552-573. [CrossRef]
5. Gui Lu, Yuan-Yuan Duan, Xiao-Dong Wang, Duu-Jong Lee. 2011. Internal flow in evaporating droplet
on heated solid surface. International Journal of Heat and Mass Transfer 54, 4437-4447. [CrossRef]
6. Alexander Alexeev, Alexander Oron. 2007. Suppression of the Rayleigh-Taylor instability of thin liquid
films by the Marangoni effect. Physics of Fluids 19, 082101. [CrossRef]

You might also like